Nothing Special   »   [go: up one dir, main page]

You seem to have javascript disabled. Please note that many of the page functionalities won't work as expected without javascript enabled.
 
 
Sign in to use this feature.

Years

Between: -

Subjects

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Journals

Article Types

Countries / Regions

remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline
remove_circle_outline

Search Results (426)

Search Parameters:
Keywords = methane adsorption

Order results
Result details
Results per page
Select all
Export citation of selected articles as:
14 pages, 8918 KiB  
Article
Molecular Energy of Metamorphic Coal and Methane Adsorption Based on Gaussian Simulation
by Tao Yang, Jingyan Hu, Tao Li, Heng Min and Shuchao Zhang
Processes 2024, 12(12), 2621; https://doi.org/10.3390/pr12122621 - 21 Nov 2024
Viewed by 232
Abstract
Effectively controlling the adsorption and desorption of coal and mine gas is crucial to preventing harm to the environment. Therefore, this paper investigated the adsorption of coal and methane molecules from the perspective of microscopic energy through Gaussian simulation. Gaussian 09W and GaussView [...] Read more.
Effectively controlling the adsorption and desorption of coal and mine gas is crucial to preventing harm to the environment. Therefore, this paper investigated the adsorption of coal and methane molecules from the perspective of microscopic energy through Gaussian simulation. Gaussian 09W and GaussView 5.0 software were used to construct and optimize the molecular model of four different metamorphic coals, namely lignite, sub-bituminous coal, bituminous coal, and anthracite, and their adsorption structure with methane as well as the energy, bond length, vibration frequency, infrared spectrum, and other data on the optimal structure were obtained. The binding energy of coal molecules and methane from large to small was as follows: sub-bituminous coal (7.3696 KJ/mol), lignite (6.6149 KJ/mol), bituminous coal (5.2170 KJ/mol), and anthracite (4.9510 KJ/mol). The equilibrium distance was negatively correlated with the binding energy, and the molecular structure and position of coal largely determined the binding energy. Additionally, adsorption was more likely to occur between methane molecules and hydroxyl groups. Many new vibration modes were observed during the adsorption of coal and methane molecules. This paper is of practical significance, as studying the adsorption of coal and mine gas can prevent and control mine gas outbursts and ensure safe production. Full article
(This article belongs to the Topic Energy Extraction and Processing Science)
Show Figures

Figure 1

Figure 1
<p>Gaussian simulation operation interface. (<b>a</b>) Energy change trend. (<b>b</b>) Vibration frequency.</p>
Full article ">Figure 2
<p>Molecular structure diagram of different metamorphic coals: (<b>a</b>) lignite (<b>b</b>) sub-bituminous coal, (<b>c</b>) bituminous coal, (<b>d</b>) anthracite.</p>
Full article ">Figure 3
<p>Optimal structure model of different metamorphic coal molecules: (<b>a</b>) lignite (<b>b</b>) sub-bituminous coal, (<b>c</b>) bituminous coal, (<b>d</b>) anthracite.</p>
Full article ">Figure 4
<p>Molecular energy values of different metamorphic coals.</p>
Full article ">Figure 5
<p>Infrared spectra of different metamorphic coal molecules. (<b>a</b>) Lignite. (<b>b</b>) Sub-bituminous coal. (<b>c</b>) Bituminous coal. (<b>d</b>) Anthracite.</p>
Full article ">Figure 5 Cont.
<p>Infrared spectra of different metamorphic coal molecules. (<b>a</b>) Lignite. (<b>b</b>) Sub-bituminous coal. (<b>c</b>) Bituminous coal. (<b>d</b>) Anthracite.</p>
Full article ">Figure 6
<p>Binding energy of different adsorption positions of lignite.</p>
Full article ">Figure 7
<p>Binding energy of different adsorption positions of sub-bituminous coal.</p>
Full article ">Figure 8
<p>Binding energy of different adsorption positions of bituminous coal.</p>
Full article ">Figure 9
<p>Binding energy of different adsorption positions of anthracite.</p>
Full article ">Figure 10
<p>Optimal structure model of methane molecules adsorbed by different metamorphic coal molecules. (<b>a</b>) lignite, (<b>b</b>) sub-bituminous coal, (<b>c</b>) bituminous coal, (<b>d</b>) anthracite.</p>
Full article ">Figure 11
<p>Adsorption equilibrium distance of different metamorphic coal molecules.</p>
Full article ">Figure 12
<p>Binding energy of different metamorphic coal molecules after adsorption.</p>
Full article ">Figure 13
<p>Infrared spectra of the optimal configuration of methane adsorbed by different metamorphic coal molecules. (<b>a</b>) Lignite. (<b>b</b>) Sub-bituminous coal. (<b>c</b>) Bituminous coal. (<b>d</b>) Anthracite.</p>
Full article ">
16 pages, 3607 KiB  
Article
Integration of CO2 Adsorbent with Ni-Al2O3 Catalysts for Enhanced Methane Production in Carbon Capture and Methanation: Cooperative Interaction of CO2 Spillover and Heat Exchange
by Dong Seop Choi, Hye Jin Kim, Jiyull Kim, Hyeona Yu and Ji Bong Joo
Catalysts 2024, 14(11), 834; https://doi.org/10.3390/catal14110834 - 20 Nov 2024
Viewed by 251
Abstract
In this study, we conducted a comparative analysis of the catalytic behavior of Ni-CaO-Al2O3 dual functional material (DFM) and a physical mixture of Ni-Al2O3 and CaO-Al2O3 in the integrated carbon capture methanation (ICCM) process [...] Read more.
In this study, we conducted a comparative analysis of the catalytic behavior of Ni-CaO-Al2O3 dual functional material (DFM) and a physical mixture of Ni-Al2O3 and CaO-Al2O3 in the integrated carbon capture methanation (ICCM) process for promoted methane production. H2-temperature-programmed surface reaction (H2-TPSR) analysis revealed that in Ni-CaO-Al2O3 DFM, CO2 adsorbed on the CaO surface can spillover to metallic Ni surface, enabling direct hydrogenation without desorption of CO2. Ni-CaO-Al2O3 DFM exhibited a rapid initial methanation rate due to CO2 spillover. The Ni-CaO-Al2O3 DFM, with Ni and CO2 adsorption sites in close distance, allows efficient utilization of the heat generated by methanation to desorb strongly adsorbed CO2, leading to enhanced methane production. Consequently, Ni-CaO-Al2O3 DFM produced 1.3 mmol/gNi of methane at 300 °C, converting 35% of the adsorbed CO2 to methane. Full article
(This article belongs to the Section Nanostructured Catalysts)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD (X-ray diffraction) patterns of catalyst and adsorbent (<b>a</b>) after calcination and (<b>b</b>) after reduction.</p>
Full article ">Figure 2
<p>CO<sub>2</sub>-TPD (CO<sub>2</sub>-temperature-programmed desorption) profiles of catalysts and adsorbents (the dotted line represents the temperature of samples).</p>
Full article ">Figure 3
<p>H<sub>2</sub>-TPSR (H<sub>2</sub>-temperature-programmed surface reaction) profiles of adsorbents and catalysts mixture.</p>
Full article ">Figure 4
<p>Thermodynamic equilibrium analysis results of the CO<sub>2</sub> methanation reaction: (<b>a</b>) product equilibrium amount and (<b>b</b>) equilibrium CO<sub>2</sub> conversion and product yield.</p>
Full article ">Figure 5
<p>The effect of temperature on CO<sub>2</sub> methanation: (<b>a</b>) CO<sub>2</sub> conversion, (<b>b</b>) CH<sub>4</sub> yield, and (<b>c</b>) CO yield, with thermodynamic equilibrium shown as gray dashed lines.</p>
Full article ">Figure 6
<p>Integrated carbon capture methanation (ICCM) results by catalyst bed configuration: (<b>a</b>) schematic diagram of catalyst bed configuration, (<b>b</b>) CO<sub>2</sub> adsorption capacity, amount of produced CH<sub>4</sub> and desorbed CO<sub>2</sub> (The symbol <span class="html-fig-inline" id="catalysts-14-00834-i001"><img alt="Catalysts 14 00834 i001" src="/catalysts/catalysts-14-00834/article_deploy/html/images/catalysts-14-00834-i001.png"/></span> represents CO<sub>2</sub> capture capacity (mmol/g<sub>bed</sub>)), (<b>c</b>) CO<sub>2</sub> conversion, (<b>d</b>) CH<sub>4</sub> production profiles in methanation step, and (<b>e</b>) cumulative CH<sub>4</sub> production.</p>
Full article ">Figure 7
<p>Integrated carbon capture methanation (ICCM) results of Ni-CaO-Al<sub>2</sub>O<sub>3</sub> DFM by operation temperature: (<b>a</b>) CO<sub>2</sub> adsorption capacities, amount of produced CH<sub>4</sub> and desorbed CO<sub>2</sub> (The symbol <span class="html-fig-inline" id="catalysts-14-00834-i001"><img alt="Catalysts 14 00834 i001" src="/catalysts/catalysts-14-00834/article_deploy/html/images/catalysts-14-00834-i001.png"/></span> represents CO<sub>2</sub> capture capacity (mmol/g<sub>bed</sub>)) and (<b>b</b>) CO<sub>2</sub> conversion.</p>
Full article ">Figure 8
<p>Schematic diagram of hydrogenation reaction pathway for (<b>a</b>) a physical mixture of catalyst and adsorbent and (<b>b</b>) the dual function material.</p>
Full article ">
17 pages, 3965 KiB  
Article
Investigation into Enhancing Methane Recovery and Sequestration Mechanism in Deep Coal Seams by CO2 Injection
by Xiongwei Sun, Hongya Wang, Bin Gong, Heng Zhao, Haoqiang Wu, Nan Wu, Wei Sun, Shizhao Zhang and Ke Jiang
Energies 2024, 17(22), 5659; https://doi.org/10.3390/en17225659 - 13 Nov 2024
Viewed by 424
Abstract
Injecting CO2 into coal seams to enhance coal bed methane (ECBM) recovery has been identified as a viable method for increasing methane extraction. This process also has significant potential for sequestering large volumes of CO2, thereby reducing the concentration of [...] Read more.
Injecting CO2 into coal seams to enhance coal bed methane (ECBM) recovery has been identified as a viable method for increasing methane extraction. This process also has significant potential for sequestering large volumes of CO2, thereby reducing the concentration of greenhouse gases in the atmosphere. However, for deep coal seams where formation pressure is relatively high, there is limited research on CO2 injection into systems with higher methane adsorption equilibrium pressure. Existing studies, mostly confined to the low-pressure stage, fail to effectively reveal the impact of factors such as temperature, high-pressure CO2 injection, and coal types on enhancing the recovery and sequestration of CO2-displaced methane. Thus, this study aims to investigate the influence of temperature, pressure, and coal types on ECBM recovery and CO2 sequestration in deep coal seams. A series of CO2 core flooding tests were conducted on various coal cores, with CO2 injection pressures ranging from 8 to 18 MPa. The CO2 and methane adsorption rates, as well as methane displacement efficiency, were calculated and recorded to facilitate result interpretation. Based on the results of these physical experiments, numerical simulation was conducted to study multi-component competitive adsorption, desorption, and seepage flow under high temperature and high pressure in a deep coal seam’s horizontal well. Finally, the optimization of the total injection amount (0.7 PV) and injection pressure (approximately 15.0 MPa) was carried out for the plan of CO2 displacement of methane in a single well in the later stage. Full article
(This article belongs to the Special Issue CO2 Capture, Utilization and Storage)
Show Figures

Figure 1

Figure 1
<p>Four coal samples: (<b>a</b>) Coal 1; (<b>b</b>) Coal 2; (<b>c</b>) Coal 3; and (<b>d</b>) Coal 4.</p>
Full article ">Figure 2
<p>Schematic of experimental set-up for core flooding experiments.</p>
Full article ">Figure 3
<p>Adsorption capacity of methane and CO<sub>2</sub> at different temperatures: (<b>a</b>) Core 1; (<b>b</b>) Core 2; (<b>c</b>) Core 3; and (<b>b</b>) Core 4.</p>
Full article ">Figure 4
<p>Maximum adsorption capacity of methane and CO<sub>2</sub> at different temperatures: (<b>a</b>) Methane and (<b>b</b>) CO<sub>2</sub>.</p>
Full article ">Figure 5
<p>The composition of the amount of methane and CO<sub>2</sub> substances in the adsorption phase: (<b>a</b>) Core 1; (<b>b</b>) Core 2; and (<b>c</b>) Core 3.</p>
Full article ">Figure 6
<p>Results of methane recovery by CO<sub>2</sub> displacement under different pressures obtained from Core 1: (<b>a</b>) 35 °C; (<b>b</b>) 45 °C; and (<b>c</b>) 55 °C.</p>
Full article ">Figure 7
<p>Results of methane recovery by CO<sub>2</sub> displacement under different temperatures. (<b>a</b>) Methane adsorption capacity after displacement; (<b>b</b>) CO<sub>2</sub> adsorption capacity after displacement; (<b>c</b>) methane desorption capacity after displacement; and (<b>d</b>) methane recovery rate.</p>
Full article ">Figure 8
<p>Simulation model of horizontal well in deep coal seam. (<b>a</b>) Daning-Jixian deep coal seam. (<b>b</b>) Simulation model.</p>
Full article ">Figure 9
<p>History matching of the typical well. (<b>a</b>) Gas rate; (<b>b</b>) Water rate; (<b>c</b>) Bottomhole pressure; (<b>d</b>) CO<sub>2</sub> molar fraction.</p>
Full article ">Figure 10
<p>The effect of displacing methane with different total amounts of CO<sub>2</sub> injection.</p>
Full article ">Figure 11
<p>The degree of improvement in methane recovery rate under different total amounts of CO<sub>2</sub> injection.</p>
Full article ">Figure 12
<p>The degree of improvement in methane recovery rate under different total amounts of CO<sub>2</sub> injection.</p>
Full article ">
13 pages, 2366 KiB  
Article
Numerical Simulation of the Coal Measure Gas Accumulation Process in Well Z-7 in Qinshui Basin
by Gaoyuan Yan, Yu Song, Fangkai Quan, Qiangqiang Cheng and Peng Wu
Processes 2024, 12(11), 2491; https://doi.org/10.3390/pr12112491 - 9 Nov 2024
Viewed by 449
Abstract
The process of coal measure gas accumulation is relatively complex, involving multiple physicochemical processes such as migration, adsorption, desorption, and seepage of multiphase fluids (e.g., methane and water) in coal measure strata. This process is constrained by multiple factors, including geological structure, reservoir [...] Read more.
The process of coal measure gas accumulation is relatively complex, involving multiple physicochemical processes such as migration, adsorption, desorption, and seepage of multiphase fluids (e.g., methane and water) in coal measure strata. This process is constrained by multiple factors, including geological structure, reservoir physical properties, fluid pressure, and temperature. This study used Well Z-7 in the Qinshui Basin as the research object as well as numerical simulations to reveal the processes of methane generation, migration, accumulation, and dissipation in the geological history. The results indicate that the gas content of the reservoir was basically zero in the early stage (before 25 Ma), and the gas content peaks all appeared after the peak of hydrocarbon generation (after 208 Ma). During the peak gas generation stage, the gas content increased sharply in the early stages. In the later stage, because of the pressurization of the hydrocarbon generation, the caprock broke through and was lost, and the gas content decreased in a zigzag manner. The reservoirs in the middle and upper parts of the coal measure were easily charged, which was consistent with the upward trend of diffusion and dissipation and had a certain relationship with the cumulative breakout and seepage dissipation. The gas contents of coal, shale, and tight sandstone reservoirs were positively correlated with the mature hydrocarbon generation of organic matter in coal seams, with the differences between different reservoirs gradually narrowing over time. Full article
Show Figures

Figure 1

Figure 1
<p>Coal stratum tectonics—burial history evolution map of the study area.</p>
Full article ">Figure 2
<p>Curve of thermal history and maturity history in the study area.</p>
Full article ">Figure 3
<p>Cumulative gas generation and gas content evolution curves of representative coal seams.</p>
Full article ">Figure 4
<p>Evolution of cumulative caprock breakthrough loss intensity and cumulative seepage loss intensity.</p>
Full article ">Figure 5
<p>Evolution of cumulative diffusion loss intensity.</p>
Full article ">Figure 6
<p>Time slice diagram of coal measure gas accumulation and evolution in Well Z-7.</p>
Full article ">
24 pages, 22042 KiB  
Article
Characterisation of the Full Pore Size Distribution of and Factors Influencing Deep Coal Reservoirs: A Case Study of the Benxi Formation in the Daning–Jixian Block at the Southeastern Margin of the Ordos Basin
by Xiaoming Chen, Tao Wang, Song Wu, Ze Deng, Julu Li, Zhicheng Ren, Daojun Huang, Wentian Fan and Gengen Zhu
Processes 2024, 12(11), 2364; https://doi.org/10.3390/pr12112364 - 28 Oct 2024
Viewed by 558
Abstract
The complex geological environment in deep layers results in differences in the pore and fracture structures and states of coalbed methane (CBM) occurrences between deep and shallow coal reservoirs. The coexistence of multiphase gases endows deep CBM with both “conventional” and “unconventional” geological [...] Read more.
The complex geological environment in deep layers results in differences in the pore and fracture structures and states of coalbed methane (CBM) occurrences between deep and shallow coal reservoirs. The coexistence of multiphase gases endows deep CBM with both “conventional” and “unconventional” geological attributes. Based on systematically collected coal samples from the Benxi Formation in the Daning–Jixian area of the Ordos Basin, high-pressure mercury intrusion (HPMI), low-temperature N2 adsorption (LTN2A), and low-pressure CO2 adsorption (LPCO2A) experiments were conducted to characterise the pore structures across the full pore size distribution of the Benxi Formation coals. The aim of this research is to gain an in-depth understanding of the pore size distribution of full-size pores and to explore the factors influencing their pore structure and control over the gas content in coal reservoirs. The results indicate that the pore size distribution of the coal samples from the Benxi Formation in the study area is unimodal and that nanopores are present. The pore sizes are relatively small, with an average total pore volume (PV) of 0.073 cm3/g and an average total specific surface area (SSA) of 227.87 m2/g. Among these, micropores account for 92.26% of the total PV and 99.57% of the total SSA, making micropores the primary contributors to the gas storage space in the Benxi Formation coals. Mesopores and macropores contribute relatively little to the PV and SSA, which is unfavourable for CBM permeability. The development of pores in the Benxi Formation coals in the study area is influenced by the coal maturity, vitrinite content, and ash yield. Generally, the PV increases when the coal’s rank increases; an increase in the vitrinite content promotes the development of micropores, whereas a relatively high ash yield leads to decreases in the PV and SSA. The influence of the SSAs of coal pores on the gas content is reflected mainly by its effect on the adsorbed gas content. Since adsorbed gas molecules exist mainly in coal pores in the adsorbed state, the SSAs of coal pores strongly affect the storage capacity of coal for adsorbed gas. Full article
Show Figures

Figure 1

Figure 1
<p>Regional location and coal-bearing strata of the Daning–Jixian block. (<b>a</b>) Location of the study area; (<b>b</b>) Tectonic location of the Daning-Jixian block; (<b>c</b>) General stratigraphic column.</p>
Full article ">Figure 2
<p>Mercury intrusion-extrusion curves of the coal samples.</p>
Full article ">Figure 3
<p>PV distribution curves based on HPMI.</p>
Full article ">Figure 4
<p>SSA distribution curves based on HPMI.</p>
Full article ">Figure 5
<p>N<sub>2</sub> adsorption-desorption isotherms.</p>
Full article ">Figure 6
<p>PV distribution curves based on LTN<sub>2</sub>A experiments.</p>
Full article ">Figure 7
<p>SSA distribution curves based on LTN<sub>2</sub>A experiments.</p>
Full article ">Figure 8
<p>CO<sub>2</sub> adsorption isotherms.</p>
Full article ">Figure 9
<p>PV distribution curves from the LPCO<sub>2</sub>A experiments.</p>
Full article ">Figure 10
<p>SSA distribution curves based on LPCO<sub>2</sub>A experiments.</p>
Full article ">Figure 11
<p>CH<sub>4</sub> adsorption isotherms.</p>
Full article ">Figure 12
<p>Principle of dominant pore size segments for LPCO<sub>2</sub>A, LTN<sub>2</sub>A, and HPMI [<a href="#B58-processes-12-02364" class="html-bibr">58</a>].</p>
Full article ">Figure 13
<p>PV distributions of different apertures in the three experiments.</p>
Full article ">Figure 14
<p>SSA distributions of different apertures combined with the three experiments.</p>
Full article ">Figure 15
<p>PV (<b>a</b>) and SSA (<b>b</b>) ratios for different pore sizes.</p>
Full article ">Figure 16
<p>Relationships between <span class="html-italic">R<sub>o,max</sub></span> and total PV and SSA (<b>a</b>), micropore PV and SSA (<b>b</b>), mesopore PV and SSA (<b>c</b>), and macropore PV and SSA (<b>d</b>).</p>
Full article ">Figure 17
<p>Relationships between vitrinite content and total PV and SSA (<b>a</b>), micropore PV and SSA (<b>b</b>), mesopore PV and SSA (<b>c</b>), and macropore PV and SSA (<b>d</b>).</p>
Full article ">Figure 18
<p>Relationships between inertinite content and total PV and SSA (<b>a</b>), micropore PV and SSA (<b>b</b>), mesopore PV and SSA (<b>c</b>), and macropore PV and SSA (<b>d</b>).</p>
Full article ">Figure 19
<p>Relationships between <span class="html-italic">A<sub>d</sub></span> and total PV and SSA (<b>a</b>), micropore PV and SSA (<b>b</b>), mesopore PV and SSA (<b>c</b>), and macropore PV and SSA (<b>d</b>).</p>
Full article ">Figure 20
<p>Relationships between the <span class="html-italic">V<sub>L</sub></span> and total PV and micropore PV (<b>a</b>), mesopore PV and macropore PV (<b>b</b>), total SSA and micropore SSA (<b>c</b>), and mesopore SSA and macropore SSA (<b>d</b>).</p>
Full article ">Figure 21
<p>Relationships between the <span class="html-italic">P<sub>L</sub></span> and total PV and micropore PV (<b>a</b>), mesopore PV and macropore PV (<b>b</b>), total SSA and micropore SSA (<b>c</b>), and mesopore SSA and macropore SSA (<b>d</b>).</p>
Full article ">
27 pages, 6198 KiB  
Article
Pd-Co Supported on Anodized Aluminium for VOCs Abatement: Reaction Mechanism, Kinetics and Applicability as Monolithic Catalyst
by Anton Naydenov, Silviya Todorova, Boriana Tzaneva, Ellie Uzunova, Hristo Kolev, Yordanka Karakirova, Daniela Karashanova and Ralitsa Velinova
Catalysts 2024, 14(10), 736; https://doi.org/10.3390/catal14100736 - 20 Oct 2024
Viewed by 920
Abstract
It has been found out that Pd-Co-based catalyst, supported on anodized aluminum, possesses very high activity in combustion reactions of C1–C6 alkanes and toluene. The catalyst characterization has been made by N2-pysisorption, XRD, SEM, XPS, FTIR, TEM, and [...] Read more.
It has been found out that Pd-Co-based catalyst, supported on anodized aluminum, possesses very high activity in combustion reactions of C1–C6 alkanes and toluene. The catalyst characterization has been made by N2-pysisorption, XRD, SEM, XPS, FTIR, TEM, and EPR methods. In view of the great interest, methane combustion was investigated in detail. It is ascertained that the complete oxidation of methane proceeds by dissociative adsorption on PdO and formation of hydroxyl and methyl groups, the former being highly reactive, and it undergoes further reaction to oxygen-containing intermediates, whereupon HCHO is one of them. The presence of Co2+ cations promotes greatly oxygen adsorption. The dissociative adsorption is favored on neighboring Co2+ cations, leading to the formation of bridging peroxides. Further, the oxygen dissociates on the nearest Pd2+ cations. According to the results from the experimental data, instrumental methods, and the observed kinetics and DFT model calculations, it can be concluded that the reaction pathway over Pd+Co/anodic alumina support (AAS) catalyst proceeds most probably through Mars–van Krevelen. The obtained data on the kinetics were used for simulation of the methane combustion in a full-scale adiabatic reactor. Full article
(This article belongs to the Special Issue Featured Papers in “Environmental Catalysis” Section)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Nitrogen physisorption isotherm (<b>A</b>) and pore size distribution (<b>B</b>) of Pd+Co/AAS.</p>
Full article ">Figure 2
<p>SEM micrographs of Pd+Co/AAS system after catalytic tests: (<b>A</b>) top view; (<b>B</b>) cross section.</p>
Full article ">Figure 3
<p>Catalytic activity of Pd+Co/AAS system in C<sub>1</sub>–C<sub>6</sub> alkanes and toluene combustion, single monolithic channel experiments under isothermal conditions.</p>
Full article ">Figure 4
<p>Depletive oxidation of methane on Pd+Co/AAS catalyst.</p>
Full article ">Figure 5
<p>XRD pattern of fresh Pd+Co/AAS catalyst.</p>
Full article ">Figure 6
<p>The FTIR spectra of anodic alumina support (AAS) and Pd+Co/AAS fresh and aged Pd+Co/AAS.</p>
Full article ">Figure 7
<p>X–ray photoelectron spectra of Pd3d, Co2p, and Al2p for Pd+Co/AAS fresh, used and aged.</p>
Full article ">Figure 8
<p>Bright Field TEM micrograph (<b>A</b>), the corresponding SAED pattern (<b>B</b>), and HRTEM (<b>C</b>) of the sample Pd+Co/AAS. The size distribution histogram of the Pd+Co nanoparticles is also presented.</p>
Full article ">Figure 9
<p>Bright Field TEM micrograph (<b>A</b>), the corresponding SAED pattern (<b>B</b>), and HRTEM (<b>C</b>) of the sample Pd/AAS. The size distribution histogram of the Pd nanoparticles is also presented.</p>
Full article ">Figure 10
<p>EPR spectra of Pd+Co/AAS (<b>A</b>) and Pd/AAS (<b>B</b>) samples.</p>
Full article ">Figure 11
<p>(<b>a</b>)—Energy diagram of the dissociative adsorption of methane on PdO(100) and on PdO(101). (<b>b</b>)—Energy diagram of oxygen adsorption on PdO(100) surface in presence of hydroxyl groups (red line) and on a CoPd15O16 cluster (blue line). The activation barriers are calculated for the ONIOM clusters.</p>
Full article ">Figure 12
<p>Dissociative adsorption of methane on Pd<sub>8</sub>O<sub>8</sub> and molecular dynamics simulation of the methane activation in presence of oxygen. Pd centers are dark-green, oxygen centers are blue, carbon is grey, and hydrogens are small light-grey balls.</p>
Full article ">Figure 13
<p>Dissociative adsorption of methane on Co<sub>2</sub>Pd<sub>6</sub>O<sub>8</sub> and molecular dynamics simulation of the methane activation in presence of oxygen. Pd centers are dark-green, oxygen centers are blue, cobalt is pink, carbon is grey, and hydrogens are small light-grey balls.</p>
Full article ">Figure 14
<p>Oxygen dissociation in the reoxidation step of Mars–van Krevelen mechanism. The arrows denote magnetic moments and their orientation. Pd centers are dark-green, oxygen centers are blue, cobalt is pink, carbon is grey, and hydrogens are small light-grey balls.</p>
Full article ">Figure 15
<p>Comparison between the experimentally measured conversions at different conditions and the model prediction by Mars–van Krevelen–mechanism, Langmuir–Hinshelwood (LH) mechanism, Eley−Rideal (ER) mechanism, and power law kinetic model (PWL).</p>
Full article ">Figure 16
<p>Model of catalytic element under semi-adiabatic conditions.</p>
Full article ">Figure 17
<p>Monolayer of PdO with the selection of ONIOM active layer used in calculations with the B3LYP density functional. Pd centers are dark-green, and oxygen centers are blue.</p>
Full article ">
20 pages, 5958 KiB  
Article
Dry Reforming of Methane (DRM) over Hydrotalcite-Based Ni-Ga/(Mg, Al)Ox Catalysts: Tailoring Ga Content for Improved Stability
by Ahmed Y. Elnour, Ahmed E. Abasaeed, Anis H. Fakeeha, Ahmed A. Ibrahim, Salwa B. Alreshaidan and Ahmed S. Al-Fatesh
Catalysts 2024, 14(10), 721; https://doi.org/10.3390/catal14100721 - 16 Oct 2024
Viewed by 739
Abstract
Dry reforming of methane (DRM) is a promising way to convert methane and carbon dioxide into syngas, which can be further utilized to synthesize value-added chemicals. One of the main challenges for the DRM process is finding catalysts that are highly active and [...] Read more.
Dry reforming of methane (DRM) is a promising way to convert methane and carbon dioxide into syngas, which can be further utilized to synthesize value-added chemicals. One of the main challenges for the DRM process is finding catalysts that are highly active and stable. This study explores the potential use of Ni-based catalysts modified by Ga. Different Ni-Ga/(Mg, Al)Ox catalysts, with various Ga/Ni molar ratios (0, 0.1, 0.3, 0.5, and 1), were synthesized by the co-precipitation method. The catalysts were tested for the DRM reaction to evaluate their activity and stability. The Ni/(Mg, Al)Ox and its Ga-modified Ni-Ga/(Mg, Al)Ox were characterized by N2 adsorption–desorption, Fourier Transform Infrared Spectroscopy (FTIR), H2-temperature-programmed reduction (TPR), X-ray diffraction (XRD), thermogravimetric analysis (TGA) and Raman techniques. The test of catalytic activity, at 700 °C, 1 atm, GHSV of 42,000 mL/h/g, and a CH4: CO2 ratio of 1, revealed that Ga incorporation effectively enhanced the catalyst stability. Particularly, the Ni-Ga/(Mg, Al)Ox catalyst with Ga/Ni ratio of 0.3 exhibited the best catalytic performance, with CH4 and CO2 conversions of 66% and 74%, respectively, and an H2/CO ratio of 0.92. Furthermore, the CH4 and CO2 conversions increased from 34% and 46%, respectively, when testing at 600 °C, to 94% and 96% when the catalytic activity was operated at 850 °C. The best catalyst’s 20 h stream performance demonstrated its great stability. DFT analysis revealed an alteration in the electronic properties of nickel upon Ga incorporation, the d-band center of the Ga modified catalyst (Ga/Ni ratio of 0.3) shifted closer to the Fermi level, and a charge transfer from Ga to Ni atoms was observed. This research provides valuable insights into the development of Ga-modified catalysts and emphasizes their potential for efficient conversion of greenhouse gases into syngas. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>FTIR spectra of the as-prepared hydrotalcites: (a) Ni-Mg-Al, (b) Ni-Ga-Mg-Al (Ga/Ni = 0.1), (c) Ni-Ga-Mg-Al (Ga/Ni = 0.3), (d) Ni-Ga-Mg-Al (Ga/Ni = 0.5), and (e) Ni-Ga-Mg-Al (Ga/Ni = 1.0).</p>
Full article ">Figure 2
<p>XRD of the as-prepared hydrotalcites: (a) Ni-Mg-Al, (b) Ni-Ga-Mg-Al (Ga/Ni = 0.1), (c) Ni-Ga-Mg-Al (Ga/Ni = 0.3), (d) Ni-Ga-Mg-Al (Ga/Ni = 0.5), and (e) Ni-Ga-Mg-Al (Ga/Ni = 1.0).</p>
Full article ">Figure 3
<p>(<b>A</b>) N<sub>2</sub> adsorption-desorption isotherms: adsorption (filled); desorption (half filled) symbols) and (<b>B</b>) pore size distribution profiles of (a) Ni/(Mg, Al)O<sub>x</sub>, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0).</p>
Full article ">Figure 4
<p>H<sub>2</sub>-TPR profiles of (<b>A</b>) hydrotalcite and (<b>B</b>) calcined samples: (a) Ni-Mg-Al, (b) Ni-Ga-Mg-Al (Ga/Ni = 0.1), (c) Ni-Ga-Mg-Al (Ga/Ni = 0.3), (d) Ni-Ga-Mg-Al (Ga/Ni = 0.5), and (e) Ni-Ga-Mg-Al (Ga/Ni = 1.0).</p>
Full article ">Figure 4 Cont.
<p>H<sub>2</sub>-TPR profiles of (<b>A</b>) hydrotalcite and (<b>B</b>) calcined samples: (a) Ni-Mg-Al, (b) Ni-Ga-Mg-Al (Ga/Ni = 0.1), (c) Ni-Ga-Mg-Al (Ga/Ni = 0.3), (d) Ni-Ga-Mg-Al (Ga/Ni = 0.5), and (e) Ni-Ga-Mg-Al (Ga/Ni = 1.0).</p>
Full article ">Figure 5
<p>XRD patterns of calcined (<b>A</b>) and reduced catalyst samples (<b>B</b>): (a) Ni/(Mg, Al)Ox, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0). ▲ Ni, ★ Ni-Ga alloy, ● Mg (GaAl)O periclase-like, ♦ NiO, ■ hydrotalcite.</p>
Full article ">Figure 6
<p>Catalytic activity tests for (<b>A</b>) CH<sub>4</sub> conversion, (<b>B</b>) CO<sub>2</sub> conversion at 700 °C, 1 atm, and GHSV of 42,000 mL/h/g for (a) Ni/(Mg, Al)Ox, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0).</p>
Full article ">Figure 6 Cont.
<p>Catalytic activity tests for (<b>A</b>) CH<sub>4</sub> conversion, (<b>B</b>) CO<sub>2</sub> conversion at 700 °C, 1 atm, and GHSV of 42,000 mL/h/g for (a) Ni/(Mg, Al)Ox, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0).</p>
Full article ">Figure 7
<p>Influence of reaction temperature on the catalytic performance for Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3) (GHSV of 42,000 mL/h/g<sub>cat</sub>, CH<sub>4</sub>/CO<sub>2</sub>/N<sub>2</sub> molar ratio of 1, and 1 atm).</p>
Full article ">Figure 8
<p>(<b>A</b>) TGA analysis and (<b>B</b>) DTA of spent catalysts: (a) Ni/(Mg, Al)O<sub>x</sub>, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0).</p>
Full article ">Figure 9
<p>TPO profiles of used catalysts: (a) Ni/(Mg, Al)O<sub>x</sub>, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0).</p>
Full article ">Figure 10
<p>Raman spectra of used catalysts: (a) Ni/(Mg, Al)O<sub>x</sub>, (b) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.1), (c) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.3), (d) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 0.5), and (e) Ni-Ga/(Mg, Al)O<sub>x</sub> (Ga/Ni = 1.0).</p>
Full article ">Figure 11
<p>(<b>A</b>) Projected electronic densities of states of Ga p and s orbitals and those of the Ni d orbitals on Ni(111) and Ni<sub>3</sub>Ga(111) surfaces and (<b>B</b>) charge density distribution on Ni(111) (top) and Ni<sub>3</sub>Ga(111) (bottom) surfaces (in e bohr<sup>−3</sup>) and their charge density difference (center); (from blue to red indicates the transition from electron depletion to accumulation).</p>
Full article ">
15 pages, 1964 KiB  
Article
On the Use of the Multi-Site Langmuir Model for Predicting Methane Adsorption on Shale
by Zhe Wu, Yuan Ji, Ke Zhang, Li Jing and Tianyi Zhao
Energies 2024, 17(19), 4990; https://doi.org/10.3390/en17194990 - 6 Oct 2024
Viewed by 726
Abstract
Shale gas, mainly consisting of adsorbed gas and free gas, has served a critical role of supplying the growing global natural gas demand in the past decades. Considering that the adsorbed methane has contributed up to 80% of the total gas in place [...] Read more.
Shale gas, mainly consisting of adsorbed gas and free gas, has served a critical role of supplying the growing global natural gas demand in the past decades. Considering that the adsorbed methane has contributed up to 80% of the total gas in place (GIP), understanding the methane adsorption behaviors is imperative to an accurate estimation of total GIP. Historically, the single-site Langmuir model, with the assumption of a homogeneous surface, is commonly applied to estimate the adsorbed gas amount. However, this assumption cannot depict the methane adsorption characteristics due to various compositions and pore sizes of shales. In this work, a multi-site model integrating the energetic heterogeneity in adsorption is derived to predict methane adsorption on shale. Our results show that the multi-site model is capable of addressing the heterogeneity of shales by a wide range of adsorption energy distributions (owing to the complex compositions and different pore sizes), which is different from the single-site model only characterized by single adsorption energy. Consequently, the multi-site model results have better accuracy against the experimental data. Therefore, applying the multi-site Langmuir model for estimating GIP in shales can achieve more accurate results compared with using the traditionally single-site model. Full article
(This article belongs to the Section H: Geo-Energy)
Show Figures

Figure 1

Figure 1
<p>Illustration of (<b>a</b>) absolute adsorption and (<b>b</b>) excess adsorption.</p>
Full article ">Figure 2
<p>Fitting results of experimental data by single-site adsorption model.</p>
Full article ">Figure 3
<p>A linear relationship between <math display="inline"><semantics> <mrow> <mi>ln</mi> <msub> <mi>P</mi> <mi>L</mi> </msub> </mrow> </semantics></math> and <math display="inline"><semantics> <mrow> <mo> </mo> <mstyle scriptlevel="0" displaystyle="true"> <mfrac> <mn>1</mn> <mi>T</mi> </mfrac> </mstyle> </mrow> </semantics></math> for four shale samples.</p>
Full article ">Figure 4
<p>Fitting results of experimental data with multi-site adsorption model.</p>
Full article ">Figure 5
<p>Comparison of multi-site model’s energy distribution with the single-site model’s constant adsorption energy for each shale sample.</p>
Full article ">Figure 6
<p>Comparison of single-site and multi-site models for the four samples.</p>
Full article ">Figure 7
<p>Comparison of absolute adsorption and excess adsorption isotherms.</p>
Full article ">
38 pages, 18101 KiB  
Review
Hydrogen Separation Membranes: A Material Perspective
by Dixit V. Bhalani and Bogyu Lim
Molecules 2024, 29(19), 4676; https://doi.org/10.3390/molecules29194676 - 1 Oct 2024
Viewed by 1250
Abstract
The global energy market is shifting toward renewable, sustainable, and low-carbon hydrogen energy due to global environmental issues, such as rising carbon dioxide emissions, climate change, and global warming. Currently, a majority of hydrogen demands are achieved by steam methane reforming and other [...] Read more.
The global energy market is shifting toward renewable, sustainable, and low-carbon hydrogen energy due to global environmental issues, such as rising carbon dioxide emissions, climate change, and global warming. Currently, a majority of hydrogen demands are achieved by steam methane reforming and other conventional processes, which, again, are very carbon-intensive methods, and the hydrogen produced by them needs to be purified prior to their application. Hence, researchers are continuously endeavoring to develop sustainable and efficient methods for hydrogen generation and purification. Membrane-based gas-separation technologies were proven to be more efficient than conventional technologies. This review explores the transition from conventional separation techniques, such as pressure swing adsorption and cryogenic distillation, to advanced membrane-based technologies with high selectivity and efficiency for hydrogen purification. Major emphasis is placed on various membrane materials and their corresponding membrane performance. First, we discuss various metal membranes, including dense, alloyed, and amorphous metal membranes, which exhibit high hydrogen solubility and selectivity. Further, various inorganic membranes, such as zeolites, silica, and CMSMs, are also discussed. Major emphasis is placed on the development of polymeric materials and membranes for the selective separation of hydrogen from CH4, CO2, and N2. In addition, cutting-edge mixed-matrix membranes are also delineated, which involve the incorporation of inorganic fillers to improve performance. This review provides a comprehensive overview of advancements in gas-separation membranes and membrane materials in terms of hydrogen selectivity, permeability, and durability in practical applications. By analyzing various conventional and advanced technologies, this review provides a comprehensive material perspective on hydrogen separation membranes, thereby endorsing hydrogen energy for a sustainable future. Full article
(This article belongs to the Section Materials Chemistry)
Show Figures

Figure 1

Figure 1
<p>Schematic representation of various raw material sources and energy sources for hydrogen production and its diversified applications.</p>
Full article ">Figure 2
<p>Schematic representation for hydrogen purification by dense metallic membranes through the solution–diffusion mechanism. Reprinted from ref. [<a href="#B55-molecules-29-04676" class="html-bibr">55</a>].</p>
Full article ">Figure 3
<p>(<b>A</b>) Digital image of stainless-steel-supported palladium membrane prepared by ELP–PP (electroless plating–pore plating) and its subfigures shows the external surface roughness measured at different areas. The measured roughness values are 3.06 ± 0.57 (<b>upper</b>), 3.44 ± 0.65 (<b>middle</b>) and 3.68 ± 0.54 μm (<b>lower</b>). Reprinted with permission from ref. [<a href="#B60-molecules-29-04676" class="html-bibr">60</a>]. (<b>B</b>) Schematic representation of supported Pd membrane.</p>
Full article ">Figure 4
<p>Schematic representation for the fabrication of zeolite LTA membrane on a functionalized support using (<b>A</b>) 1,4-diisocyanate (DIC-4) as the covalent linker, (<b>B</b>) 3-chloropropyltrimethoxysilane (CPTMS) as the covalent linker, and (<b>C</b>) 3-aminopropyltriethoxysilane (APTES) as the covalent linker. Reprinted with permission from refs. [<a href="#B88-molecules-29-04676" class="html-bibr">88</a>,<a href="#B89-molecules-29-04676" class="html-bibr">89</a>,<a href="#B90-molecules-29-04676" class="html-bibr">90</a>].</p>
Full article ">Figure 5
<p>Schematic representation of composite silica membrane.</p>
Full article ">Figure 6
<p>Preparation of stainless-steel-supported silica membranes and their SEM images. Reprinted with permission from ref. [<a href="#B101-molecules-29-04676" class="html-bibr">101</a>].</p>
Full article ">Figure 7
<p>Schematic representation of amorphous silica networks derived by TEOS (<b>a</b>) and BTESE (<b>b</b>). Reprinted with permission from ref. [<a href="#B110-molecules-29-04676" class="html-bibr">110</a>].</p>
Full article ">Figure 8
<p>Schematic representation of the pyrolysis setup. Reprinted with permission from ref. [<a href="#B132-molecules-29-04676" class="html-bibr">132</a>].</p>
Full article ">Figure 9
<p>Schematic representation of various gas transport mechanisms.</p>
Full article ">Figure 10
<p>Representation for membrane preparation by the non-solvent-induced phase-inversion process.</p>
Full article ">Figure 11
<p>Structures of various promising polymers used for the fabrication of polymeric membranes.</p>
Full article ">Figure 12
<p>Schematic representation of (<b>A</b>) H<sub>2</sub>-selective polymer membranes and (<b>B</b>) CO<sub>2</sub>-selective polymer membranes.</p>
Full article ">Figure 13
<p>Representation of PBI structural packing: π–π stacking and hydrogen bonding. Reprinted with permission from ref. [<a href="#B152-molecules-29-04676" class="html-bibr">152</a>].</p>
Full article ">Figure 14
<p>(<b>A</b>) Schematic representation of the chemical structure of polymers and hydrogen-bonding interactions between the functional groups of Matrimid and PBI. (<b>B</b>) Proposed mechanism for the chemical cross-linking modification of the Matrimid component of the blend using p-xylene diamine. (<b>C</b>) Chemical structure of p-xylene diamine and (<b>D</b>) possible chain morphology and configuration of p-xylene diamine cross-linked with Matrimid (cross-linking agents are specified by ovals). Reprinted with permission from ref. [<a href="#B153-molecules-29-04676" class="html-bibr">153</a>].</p>
Full article ">Figure 15
<p>Schematic illustration of (<b>a</b>) proton transfer mechanism and hydrogen bonding in the PBI–H<sub>3</sub>PO<sub>4</sub> complex and (<b>b</b>) the preparation of H<sub>3</sub>PO<sub>4</sub>-doped PBI films with PBI backbones cross-linked by acids. SEM images with an overlaid SEM/EDS mapping of phosphorus on the (<b>c</b>) surface and (<b>d</b>) cross-section of a PBI–(H<sub>3</sub>PO<sub>4</sub>) 1.0 film. The red dots display the distribution of phosphorus in the polymer, (<b>e</b>) the comparative permselectivity of the acid-doped PBI. Reprinted with permission from ref. [<a href="#B154-molecules-29-04676" class="html-bibr">154</a>].</p>
Full article ">Figure 16
<p>(<b>A</b>) Effects of the coagulation bath temperature on real selectivity and gas permeance. (<b>B</b>) Effects of different non-solvents on selectivity and H<sub>2</sub> permeance. (<b>C</b>) Effects of sequential coating on permeance. (<b>D</b>) Effects of sequential coating on the selectivity of hydrogen/methane binary mixture with a 50–50% concentration at 1 bar and 25 °C for film casting and dip coating. Reprinted with permission from ref. [<a href="#B155-molecules-29-04676" class="html-bibr">155</a>].</p>
Full article ">Figure 17
<p>Schematics of dual-layered hollow-fiber spinning setup and triple-orifice spinneret: (a) dope fluid tank and pump; (b) bore fluid tank and pump; (c) filter; (d) spinneret; (e) coagulation bath; and (f) take-up drum. Reprinted with permission from ref. [<a href="#B157-molecules-29-04676" class="html-bibr">157</a>].</p>
Full article ">Figure 18
<p>(<b>A</b>–<b>D</b>) Representation of various steps in the fabrication of CNT/PES membranes. (<b>E</b>) Digital images of prepared PES membranes and CNT/PES membranes. Reprinted with permission from ref. [<a href="#B166-molecules-29-04676" class="html-bibr">166</a>].</p>
Full article ">Figure 19
<p>SEM images of (<b>A</b>,<b>C</b>) PES membrane and (<b>B</b>,<b>D</b>) CNTs/PES. (<b>E</b>,<b>F</b>) Schematic representation of pore structure and gas transport mechanism through the PES pore walls and CNT/PES flat-smooth walls. Reprinted with permission from ref. [<a href="#B166-molecules-29-04676" class="html-bibr">166</a>].</p>
Full article ">Figure 20
<p>(<b>A</b>) Schematic representation of the experimental setup utilized to measure the permeability of membrane gases. (<b>B</b>) Digital image of CNT/PES membranes after a gas separation experiment. Reprinted with permission from ref. [<a href="#B166-molecules-29-04676" class="html-bibr">166</a>].</p>
Full article ">Figure 21
<p>(<b>A</b>) Three-dimensional structure of various MOF materials. Reprinted with permission from ref. [<a href="#B179-molecules-29-04676" class="html-bibr">179</a>]. (<b>B</b>) Schematic representation of porous inorganic filler-based MMMs for gas separation. Reprinted with permission from ref. [<a href="#B168-molecules-29-04676" class="html-bibr">168</a>].</p>
Full article ">Figure 22
<p>Gas-separation performance of bare PBI membranes and 20 wt.%-loaded MMMs containing ZIF-93, ZIF-11, and the ZIF-93/11 hybrid materials, which were synthesized in DMAc and MeOH. The continuous lines correspond to the original Robeson upper bounds of 1991 and 2008, and the dashed line corresponds to the upper bound calculated for 180 °C. Reprinted with permission from ref. [<a href="#B180-molecules-29-04676" class="html-bibr">180</a>].</p>
Full article ">Figure 23
<p>(<b>a</b>) ZIF-7 crystal-phase diagram as a function of concentrations of zinc and bIm. (<b>b</b>) Illustration of the ZIF-7 synthesis stages and the corresponding conditions during the PMMOF process. Reprinted with permission from ref. [<a href="#B174-molecules-29-04676" class="html-bibr">174</a>].</p>
Full article ">Figure 24
<p>(<b>A</b>) H<sub>2</sub>/CO<sub>2</sub>, H<sub>2</sub>/CH<sub>4</sub>, and H<sub>2</sub>/N<sub>2</sub> selectivity of MMMs (PG1N, PG3N, and PG5N) at 1 bar and 20, 40, and 60 °C. Reprinted with permission from ref. [<a href="#B181-molecules-29-04676" class="html-bibr">181</a>]. (<b>B</b>) Schematic representation of ZIF-8 gel in PIM-1 membrane matrix. Reprinted with permission from ref. [<a href="#B182-molecules-29-04676" class="html-bibr">182</a>].</p>
Full article ">
17 pages, 4825 KiB  
Article
Investigation into the Simulation and Mechanisms of Metal–Organic Framework Membrane for Natural Gas Dehydration
by Qingxiang Song, Pengxiao Liu, Congjian Zhang, Yao Ning, Xingjian Pi and Ying Zhang
Nanomaterials 2024, 14(19), 1583; https://doi.org/10.3390/nano14191583 - 30 Sep 2024
Viewed by 509
Abstract
Natural gas dehydration is a critical process in natural gas extraction and transportation, and the membrane separation method is the most suitable technology for gas dehydration. In this paper, based on molecular dynamics theory, we investigate the performance of a metal–organic composite membrane [...] Read more.
Natural gas dehydration is a critical process in natural gas extraction and transportation, and the membrane separation method is the most suitable technology for gas dehydration. In this paper, based on molecular dynamics theory, we investigate the performance of a metal–organic composite membrane (ZIF-90 membrane) in natural gas dehydration. The paper elucidates the adsorption, diffusion, permeation, and separation mechanisms of water and methane with the ZIF-90 membrane, and clarifies the influence of temperature on gas separation. The results show that (1) the diffusion energy barrier and pore size are the primary factors in achieving the separation of water and methane. The diffusion energy barriers for the two molecules (CH4 and H2O) are ΔE(CH4) = 155.5 meV and ΔE(H2O) = 50.1 meV, respectively. (2) The ZIF-90 is more selective of H2O, which is mainly due to the strong interaction between the H2O molecule and the polar functional groups (such as aldehyde groups) within the ZIF-90. (3) A higher temperature accelerates the gas separation process. The higher the temperature is, the faster the separation process is. (4) The pore radius is identified as the intrinsic mechanism enabling the separation of water and methane in ZIF-90 membranes. Full article
(This article belongs to the Special Issue Advanced Nanostructured Membranes)
Show Figures

Figure 1

Figure 1
<p>Structure of the ZIF-90. Purple: Zn; Red: O; Black: C; Green: N.</p>
Full article ">Figure 2
<p>The diffusion energy barrier for (<b>a</b>) H<sub>2</sub>O and (<b>b</b>) CH<sub>4</sub>.</p>
Full article ">Figure 3
<p>Adsorption isotherms for (<b>a</b>) CH<sub>4</sub> and (<b>b</b>) H<sub>2</sub>O on ZIF-90 at different temperatures.</p>
Full article ">Figure 4
<p>MSD (<b>a</b>–<b>c</b>) and diffusion coefficient (<b>d</b>) of CH<sub>4</sub> on ZIF-90 at different temperatures. (<b>a</b>) 20 CH<sub>4</sub> molecules, (<b>b</b>) 40 CH<sub>4</sub> molecules, (<b>c</b>) 60 CH<sub>4</sub> molecules. (<b>d</b>) The diffusion coefficient.</p>
Full article ">Figure 5
<p>MSD (<b>a</b>–<b>c</b>) and diffusion coefficient (<b>d</b>) of H<sub>2</sub>O on ZIF-90 at different temperatures. (<b>a</b>) 20 H<sub>2</sub>O molecules, (<b>b</b>) 40 H<sub>2</sub>O molecules, (<b>c</b>) 60 H<sub>2</sub>O molecules. (<b>d</b>) The diffusion coefficient.</p>
Full article ">Figure 6
<p>Schematic diagram of the different regions of the ZIF-90.</p>
Full article ">Figure 7
<p>The number of H<sub>2</sub>O molecules in regions A, B, and C changing over time during the process of 100 H<sub>2</sub>O gas molecules permeating through the ZIF-90 membrane at different temperatures: (<b>a</b>) 300 K, (<b>b</b>) 400 K, (<b>c</b>) 500 K, (<b>d</b>) 600 K, and (<b>e</b>) 900 K.</p>
Full article ">Figure 8
<p>The number of H<sub>2</sub>O molecules in regions A, B, and C changing over time during the process of 200 H<sub>2</sub>O molecules permeating through the ZIF-90 membrane at different temperatures: (<b>a</b>) 400 K, (<b>b</b>) 500 K, (<b>c</b>) 600 K, (<b>d</b>) 900 K.</p>
Full article ">Figure 9
<p>Number of CH<sub>4</sub> molecules in regions A, B, and C changing over time during the process of 100 CH<sub>4</sub> gas molecules permeating through the ZIF-90 membrane at different temperatures: (<b>a</b>) 300 K, (<b>b</b>) 400 K, (<b>c</b>) 500 K, (<b>d</b>) 600 K, (<b>e</b>) 900 K.</p>
Full article ">Figure 10
<p>Number of CH<sub>4</sub> molecules in regions A, B, and C changing over time during the process of 200 CH<sub>4</sub> gas molecules permeating through the ZIF-90 membrane at different temperatures: (<b>a</b>) 300 K, (<b>b</b>) 400 K, (<b>c</b>) 500 K, (<b>d</b>) 600 K.</p>
Full article ">Figure 11
<p>Final state simulation snapshots of the gas mixture (100 H<sub>2</sub>O and 100 CH<sub>4</sub>) permeating through the ZIF-90 membrane at (<b>a</b>) 400 K, (<b>b</b>) 500 K, (<b>c</b>) 600 K, and (<b>d</b>) 900 K for 5 ns. (Grey: C; White: H; Red: O).</p>
Full article ">Figure 12
<p>Number of H<sub>2</sub>O molecules in regions A, B, and C changing over time during the process of the mixture gas (100 H<sub>2</sub>O and 100 CH<sub>4</sub>) permeating through the ZIF-90 membrane at different temperatures: (<b>a</b>) 400 K, (<b>b</b>) 500 K, (<b>c</b>) 600 K, (<b>d</b>) 900 K.</p>
Full article ">
21 pages, 1872 KiB  
Proceeding Paper
Recent Advances in Membrane Technologies for Biogas Upgrading
by Gabriella Aguilloso, Kimberly Arpia, Morzina Khan, Zachary Alijah Sapico and Edgar Clyde Repato Lopez
Eng. Proc. 2024, 67(1), 57; https://doi.org/10.3390/engproc2024067057 - 30 Sep 2024
Viewed by 1095
Abstract
The pressing environmental and energy challenges of today are driven by the depletion of fossil fuels and a surge in greenhouse gas emissions, particularly carbon dioxide. This situation highlights the critical need for sustainable energy solutions. While carbon capture and storage (CCS) technologies [...] Read more.
The pressing environmental and energy challenges of today are driven by the depletion of fossil fuels and a surge in greenhouse gas emissions, particularly carbon dioxide. This situation highlights the critical need for sustainable energy solutions. While carbon capture and storage (CCS) technologies offer hope, they face economic challenges at the scale needed to significantly reduce carbon dioxide emissions. Biogas, produced mainly through the anaerobic digestion of various biomass sources like agricultural waste, municipal solid waste, and wastewater, presents a renewable alternative. Composed largely of methane and carbon dioxide, biogas can be upgraded to bio-methane, serving as an eco-friendly replacement for natural gas. Technological advancements, particularly in membrane separation, have made biogas purification more efficient and cost-effective. Anaerobic digestion, a key process in biogas production, breaks down organic matter into simpler compounds, which are then transformed into gases like methane and carbon dioxide. The composition of biogas depends on the feedstock and digestion conditions, with methane being a valuable but challenging component to separate due to its greenhouse gas properties. Several purification technologies have been developed, including absorption, adsorption, cryogenic separation, and membrane separation, each with unique benefits and drawbacks. Membrane separation is particularly promising for its environmental benefits and scalability. However, the biogas industry faces challenges, especially in developing countries, due to high costs and limited research and development. Overcoming these obstacles requires collaboration among various stakeholders. Looking ahead, the future of biogas technology is bright, with advances in membrane materials and integrated refining processes. Integrating biogas into sectors like waste management and agriculture is crucial for its development and for meeting global renewable energy goals. Biogas technology not only reduces dependence on fossil fuels but also plays a vital role in the transition to sustainable energy. Full article
(This article belongs to the Proceedings of The 3rd International Electronic Conference on Processes)
Show Figures

Figure 1

Figure 1
<p>Biogas upgrading by chemical absorption of CO<sub>2</sub>. Reprinted with permission from [<a href="#B9-engproc-67-00057" class="html-bibr">9</a>]. Copyright Springer.</p>
Full article ">Figure 2
<p>Pressure Swing Adsorption. Reprinted with permission from [<a href="#B10-engproc-67-00057" class="html-bibr">10</a>]. Copyright Elsevier.</p>
Full article ">Figure 3
<p>Cryogenic separation process diagram. Reprinted with permission from [<a href="#B10-engproc-67-00057" class="html-bibr">10</a>]. Copyright Elsevier.</p>
Full article ">Figure 4
<p>Membrane separation process for biogas upgrading. Reprinted with permission from [<a href="#B10-engproc-67-00057" class="html-bibr">10</a>]. Copyright Elsevier.</p>
Full article ">Figure 5
<p>Single-stage membrane-based biogas upgrading process. Process (<b>a</b>) the permeate flows to the ambient. Process (<b>b</b>) the permeate is partially recycled to enhance the CH<sub>4</sub> recovery. Reprinted with permission from [<a href="#B5-engproc-67-00057" class="html-bibr">5</a>]. Copyright The Royal Society of Chemistry.</p>
Full article ">Figure 6
<p>Two-stage gas permeation process for biogas upgrading. Process (<b>a</b>) only needs one compressor and recycles the permeate of the second step. Process (<b>b</b>) needs two compressors and recycles the retentate of the second step. In Process (<b>c</b>), two compressors are needed, and the retentate of the second step is mixed with the one coming from the first step. In process (<b>d</b>), the feed gas is used as a sweep gas on the permeate side of the second module. Reprinted with permission from [<a href="#B5-engproc-67-00057" class="html-bibr">5</a>]. Copyright The Royal Society of Chemistry.</p>
Full article ">Figure 7
<p>Three-stage gas permeation process for biogas upgrading. Reprinted with permission from [<a href="#B5-engproc-67-00057" class="html-bibr">5</a>]. Copyright The Royal Society of Chemistry.</p>
Full article ">
22 pages, 5397 KiB  
Article
Synthesis, Characterization, and Attrition Resistance of Kaolin and Boehmite Alumina-Reinforced La0.7Sr0.3FeO3 Perovskite Catalysts for Chemical Looping Partial Oxidation of Methane
by Farzam Fotovat, Mohammad Beyzaei, Hadi Ebrahimi and Erfan Mohebolkhames
Catalysts 2024, 14(10), 670; https://doi.org/10.3390/catal14100670 - 27 Sep 2024
Viewed by 624
Abstract
This study investigates the impact of kaolin and boehmite alumina binders on the synthesis, catalytic properties, and attrition resistance of a La0.7Sr0.3FeO3 (LSF) perovskite catalyst designed for the chemical looping partial oxidation (CLPO) of methane to produce synthesis [...] Read more.
This study investigates the impact of kaolin and boehmite alumina binders on the synthesis, catalytic properties, and attrition resistance of a La0.7Sr0.3FeO3 (LSF) perovskite catalyst designed for the chemical looping partial oxidation (CLPO) of methane to produce synthesis gas sustainably. The as-synthesized and used catalysts with varying kaolin and boehmite alumina contents (KB(x,y)/LSF) were scrutinized by a variety of characterization methods, including XRD, FE-SEM/EDS, BET, TPD-NH3, and TPD-O2 techniques. The catalytic activity of the synthesized samples was tested at 800 to 900 °C in a fixed-bed reactor producing syngas through the CLPO process over the consecutive redox cycles. Additionally, the attrition resistance of the fresh and used catalyst samples was examined in a jet cup apparatus to assess their durability against the stresses induced by thermal shocks or changes in the crystal lattice caused by chemical reactions. The characterization results showed the pure perovskite crystal structure of KB(x,y)/LSF catalysts demonstrating adequate oxygen adsorption capacity, effective coke mitigation capability, robust thermal stability, and resilience to agglomeration during repetitive redox cycles. Among the tested catalysts, KB(25,15)/LSF was identified as the superior sample, as it could consistently produce syngas with a suitable H2:CO molar ratio varying from 2 to 3 within ten redox cycles at 900 °C, with CH4 conversion and CO selectivity values up to 64% and 87%, respectively. The synthesized catalysts demonstrated a logarithmic attrition pattern in the jet cup tests at room temperature, featuring high attrition resistance after the erosion of particle shape irregularities or weakly bound particles. Moreover, the KB(25,15)/LSF catalyst used at 900 °C showed great resistance in the attrition test, warranting its endurance in the face of extraordinarily harsh conditions in fluidized bed reactors employed for the CLPO process. Full article
(This article belongs to the Special Issue Fluidizable Catalysts for Novel Chemical Processes)
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>XRD patterns of (<b>a</b>) the synthesized KB/LSF oxygen carriers with different KB contents; (<b>b</b>) fresh KB<sub>(25,15)</sub>/LSF and KB<sub>(25,15)</sub>/LSF used in the fixed-bed reactor at 900 °C over 10 cycles.</p>
Full article ">Figure 2
<p>FE-SEM image and EDS analysis of (<b>a</b>,<b>b</b>) fresh KB<sub>(25,15)</sub>/LSF and (<b>c</b>,<b>d</b>) the used KB<sub>(25,15)</sub>/LSF used in the fixed-bed reactor at 900 °C over 10 cycles.</p>
Full article ">Figure 3
<p>N<sub>2</sub> adsorption/desorption isotherms for fresh KB<sub>(10,30)</sub>/LSF and KB<sub>(30,10)</sub>/LSF at −196.15 °C.</p>
Full article ">Figure 4
<p>BJH profiles of fresh and used KB<sub>(25,15)</sub>/LSF.</p>
Full article ">Figure 5
<p>Oxygen adsorption profile of the KB<sub>(25, 15)</sub>/LSF catalyst.</p>
Full article ">Figure 6
<p>TPD-NH<sub>3</sub> profile of the KB<sub>(25,15)</sub>/LSF catalyst.</p>
Full article ">Figure 7
<p>(<b>a</b>) H<sub>2</sub>, (<b>b</b>) CO, and (<b>c</b>) CO<sub>2</sub> molar concentration in the gaseous products of the tests conducted at 800, 850, and 900 °C.</p>
Full article ">Figure 7 Cont.
<p>(<b>a</b>) H<sub>2</sub>, (<b>b</b>) CO, and (<b>c</b>) CO<sub>2</sub> molar concentration in the gaseous products of the tests conducted at 800, 850, and 900 °C.</p>
Full article ">Figure 8
<p>(<b>a</b>) CH<sub>4</sub> conversion, (<b>b</b>) H<sub>2</sub>:CO molar ratio, and (<b>c</b>) CO selectivity of the tests conducted at 800, 850, and 900 °C.</p>
Full article ">Figure 8 Cont.
<p>(<b>a</b>) CH<sub>4</sub> conversion, (<b>b</b>) H<sub>2</sub>:CO molar ratio, and (<b>c</b>) CO selectivity of the tests conducted at 800, 850, and 900 °C.</p>
Full article ">Figure 9
<p>Cyclic performance of KB<sub>(25,15)</sub>/LSF in reduction phase when POM reaction prevails in ten redox cycles at 900 °C.</p>
Full article ">Figure 10
<p>Attrition curve of the synthesized LSF samples with varying kaolin and boehmite contents.</p>
Full article ">Figure 11
<p><span class="html-italic">A</span><sub>tot</sub> and <span class="html-italic">A</span><sub>i</sub> values for the synthesized LSF samples with varying kaolin and boehmite contents.</p>
Full article ">Figure 12
<p>Schematics of the fixed-bed reactor used in this study for the reactivity tests.</p>
Full article ">
17 pages, 2720 KiB  
Article
Comprehensive Characterization and Metamorphic Control Analysis of Full Apertures in Different Coal Ranks within Deep Coal Seams
by Qi Li, Yong Wu and Lei Qiao
Appl. Sci. 2024, 14(18), 8566; https://doi.org/10.3390/app14188566 - 23 Sep 2024
Viewed by 514
Abstract
The pore fracture structure of deep coal reservoirs is crucial for evaluating the potential of deep coalbed methane resources, conducting exploration and development, and controlling coal mine gas disasters. Mercury intrusion porosimetry, the liquid nitrogen method, and the low-temperature carbon dioxide adsorption method [...] Read more.
The pore fracture structure of deep coal reservoirs is crucial for evaluating the potential of deep coalbed methane resources, conducting exploration and development, and controlling coal mine gas disasters. Mercury intrusion porosimetry, the liquid nitrogen method, and the low-temperature carbon dioxide adsorption method were used to study the full pore size structure and pore fractal characteristics of different coal grades in deep coal and comprehensively characterize the pore structure of kilometer-level coal mining. The sponge, Frenkel–Halsey–Hill (FHH), and density function models were applied to comprehensively analyze the pore complexity of coal, and the influence of metamorphic degree on pore size structure was evaluated. The distribution relationship of pore volume in different stages of coal samples was macropore→mesopore→micropore, and macropores had the best connectivity. Micropores and mesopores had the largest specific surface area, and the development of micropores and microcracks controlled the deep gas adsorption performance. The micropore volume and specific surface area both revealed a nonlinear decreasing trend with the increase in volatile matter, and coal metamorphism promoted the development of micropores. The pore volume and specific surface area of mesopores and macropores decreased first and then increased in a “U” shape with increasing volatile matter. In contrast, the fractal dimension D1 revealed an inverted U shape with increasing volatile matter, followed by a decrease. The D2 value decreased nonlinearly with increasing volatile matter, whereas the D3 value increased nonlinearly with increasing volatile matter. The degree of metamorphism increased, and the microporous structure became more regular. Full article
Show Figures

Figure 1

Figure 1
<p>Mercury intrusion curves of samples from different coal ranks. (<b>a</b>) XJE, (<b>b</b>) PML, (<b>c</b>) HYS, (<b>d</b>) LXK.</p>
Full article ">Figure 2
<p>Adsorption–desorption curve of liquid nitrogen method. (<b>a</b>) XJE, (<b>b</b>) PML, (<b>c</b>) HYS, (<b>d</b>) LXK.</p>
Full article ">Figure 3
<p>Low-temperature carbon dioxide adsorption–desorption curve. (<b>a</b>) XJE, (<b>b</b>) PML, (<b>c</b>) HYS, (<b>d</b>) LXK.</p>
Full article ">Figure 4
<p>Distribution of full aperture pore volume. (<b>a</b>) Stage pore volume, (<b>b</b>) accumulated pore volume.</p>
Full article ">Figure 5
<p>Distribution of full aperture-specific surface area. (<b>a</b>) Stage surface area, (<b>b</b>) accumulated surface area.</p>
Full article ">Figure 6
<p>Relationship between volatile matter, pore volume, and specific surface area. (<b>a</b>) Pore volume, (<b>b</b>) surface area.</p>
Full article ">Figure 7
<p>Fractal calculation of sponge models with different coal ranks. (<b>a</b>) XJE, (<b>b</b>) PML, (<b>c</b>) HYS, and (<b>d</b>) LXK.</p>
Full article ">Figure 8
<p>Fractal calculation of FHH models for different coal ranks. (<b>a</b>) XJE, (<b>b</b>) PML, (<b>c</b>) HYS, and (<b>d</b>) LXK.</p>
Full article ">Figure 9
<p>Fractal calculation of density function models for different coal ranks. (<b>a</b>) XJE, (<b>b</b>) PML, (<b>c</b>) HYS, and (<b>d</b>) LXK.</p>
Full article ">Figure 10
<p>Relationship between volatile matter and fractal dimension.</p>
Full article ">
31 pages, 3833 KiB  
Article
Transition Metal-Promoted LDH-Derived CoCeMgAlO Mixed Oxides as Active Catalysts for Methane Total Oxidation
by Marius C. Stoian, Cosmin Romanitan, Katja Neubauer, Hanan Atia, Constantin Cătălin Negrilă, Ionel Popescu and Ioan-Cezar Marcu
Catalysts 2024, 14(9), 625; https://doi.org/10.3390/catal14090625 - 17 Sep 2024
Viewed by 799
Abstract
A series of M(x)CoCeMgAlO mixed oxides with different transition metals (M = Cu, Fe, Mn, and Ni) with an M content x = 3 at. %, and another series of Fe(x)CoCeMgAlO mixed oxides with Fe contents x ranging from 1 to 9 at. [...] Read more.
A series of M(x)CoCeMgAlO mixed oxides with different transition metals (M = Cu, Fe, Mn, and Ni) with an M content x = 3 at. %, and another series of Fe(x)CoCeMgAlO mixed oxides with Fe contents x ranging from 1 to 9 at. % with respect to cations, while keeping constant in both cases 40 at. % Co, 10 at. % Ce and Mg/Al atomic ratio of 3 were prepared via thermal decomposition at 750 °C in air of their corresponding layered double hydroxide (LDH) precursors obtained by coprecipitation. They were tested in a fixed bed reactor for complete methane oxidation with a gas feed of 1 vol.% methane in air to evaluate their catalytic performance. The physico-structural properties of the mixed oxide samples were investigated with several techniques, such as powder X-ray diffraction (XRD), scanning electron microscopy (SEM) coupled with energy dispersive X-ray spectroscopy (EDX), elemental mappings, inductively coupled plasma optical emission spectroscopy (ICP-OES), X-ray photoelectron spectroscopy (XPS), temperature-programmed reduction under hydrogen (H2-TPR) and nitrogen adsorption–desorption at −196 °C. XRD analysis revealed in all the samples the presence of Co3O4 crystallites together with periclase-like and CeO2 phases, with no separate M-based oxide phase. All the cations were distributed homogeneously, as suggested by EDX measurements and elemental mappings of the samples. The metal contents, determined by EDX and ICP-OES, were in accordance with the theoretical values set for the catalysts’ preparation. The redox properties studied by H2-TPR, along with the surface composition determined by XPS, provided information to elucidate the catalytic combustion properties of the studied mixed oxide materials. The methane combustion tests showed that all the M-promoted CoCeMgAlO mixed oxides were more active than the M-free counterpart, the highest promoting effect being observed for Fe as the doping transition metal. The Fe(x)CoCeMgAlO mixed oxide sample, with x = 3 at. % Fe displayed the highest catalytic activity for methane combustion with a temperature corresponding to 50% methane conversion, T50, of 489 °C, which is ca. 40 °C lower than that of the unpromoted catalyst. This was attributed to its superior redox properties and lowest activation energy among the studied catalysts, likely due to a Fe–Co–Ce synergistic interaction. In addition, long-term tests of Fe(3)CoCeMgAlO mixed oxide were performed, showing good stability over 60 h on-stream. On the other hand, the addition of water vapors in the feed led to textural and structural changes in the Fe(3)CoCeMgAlO system, affecting its catalytic performance in methane complete oxidation. At the same time, the catalyst showed relatively good recovery of its catalytic activity as soon as the water vapors were removed from the feed. Full article
Show Figures

Graphical abstract

Graphical abstract
Full article ">Figure 1
<p>Diffractograms of (<b>a</b>) M(3)CoCeMgAl, and (<b>b</b>) Fe(x)CoCeMgAl LDH-based precursors compared to that of undoped CoCeMgAl LDH. Symbols: #—LDH phase; ∗—boehmite (AlOOH) phase.</p>
Full article ">Figure 2
<p>Diffractograms of (<b>a</b>) M(3)CoCeMgAlO and (<b>b</b>) Fe(x)CoCeMgAlO mixed oxides calcined at 750 °C compared to their unpromoted CoCeMgAlO counterpart. Symbols: Δ—Co<sub>3</sub>O<sub>4</sub> phase; ∗—CeO<sub>2</sub> phase; #—Mg(Al)O phase.</p>
Full article ">Figure 3
<p>(<b>a</b>) High-resolution O 1s core level and (<b>b</b>) C 1s core level X-ray photoelectron profiles of the Fe(x)CoCeMgAlO mixed oxide samples: CoCeMgAlO (A); Fe(1)CoCeMgAlO (B); Fe(3)CoCeMgAlO (C); Fe(6)CoCeMgAlO (D); Fe(9)CoCeMgAlO (E).</p>
Full article ">Figure 4
<p>(<b>a</b>) High-resolution Co 2p core level and (<b>b</b>) Ce 3d core level X-ray photoelectron profiles of the Fe(x)CoCeMgAlO mixed oxide samples: CoCeMgAlO (A); Fe(1)CoCeMgAlO (B); Fe(3)CoCeMgAlO (C); Fe(6)CoCeMgAlO (D); Fe(9)CoCeMgAlO (E).</p>
Full article ">Figure 5
<p>High-resolution Fe 2p core level X-ray photoelectron profiles of the Fe(x)CoCeMgAlO mixed oxide catalysts: Fe(1)CoCeMgAlO (A); Fe(3)CoCeMgAlO (B); Fe(6)CoCeMgAlO (C); Fe(9)CoCeMgAlO (D).</p>
Full article ">Figure 6
<p>H<sub>2</sub>-TPR profiles of CoCeMgAlO and promoted M(3)CoCeMgAlO mixed oxides.</p>
Full article ">Figure 7
<p>H<sub>2</sub>-TPR profiles of the promoted Fe(x)CoCeMgAlO catalysts.</p>
Full article ">Figure 8
<p>The light-off curves for the methane combustion reaction over (<b>a</b>) CoCeMgAlO and M(3)CoCeMgAlO and (<b>b</b>) Fe(x)CoCeMgAlO catalysts. Reaction conditions: 1 vol.% methane in air, GHSV of 16,000 h<sup>−1</sup>, 1 cm<sup>3</sup> of catalyst.</p>
Full article ">Figure 9
<p>Variation of the total hydrogen consumption below 750 °C in the H<sub>2</sub>-TPR measurements and of the intrinsic reaction rates at 400 and 450 °C versus Fe content in the Fe(x)CoCeMgAlO series.</p>
Full article ">Figure 10
<p>(<b>a</b>) Dependence of the Ce/Co surface atomic ratio and of the intrinsic reaction rates at 400 and 450 °C on the Fe content in the Fe(x)CoCeMgAlO series. (<b>b</b>) Dependence between the intrinsic reaction rate at 400 °C and the Ce<sup>4+</sup>/Ce surface atomic ratio in the Fe(x)CoCeMgAlO series.</p>
Full article ">Figure 11
<p>The dependence of the methane total oxidation on the gas hourly space velocity (GHSV) at constant 1 vol. % methane concentration in the feed gas for the Fe(3)CoCeMgAlO catalyst.</p>
Full article ">Figure 12
<p>Evolution of methane conversion at 600 °C with time over Fe(3)CoCeMgAlO catalyst. Reaction conditions: 1 vol.% CH<sub>4</sub> in air and GHSV of 16,000 h<sup>−1</sup> with 1 cm<sup>3</sup> of catalyst.</p>
Full article ">Figure 13
<p>Evolution of methane conversion with time on stream during combustion tests at 600 °C for the Fe(3)CoCeMgAlO catalyst in dry/humid conditions runs. Dry reaction conditions: 1 vol.% CH<sub>4</sub> in air and GHSV of 16,000 h<sup>−1</sup>, 1 cm<sup>3</sup> of catalyst. Humid reaction conditions were obtained by adding, with a peristaltic pump, a flow of 0.14 mL min<sup>−1</sup> of deionized liquid water to the dry mixture, corresponding to a water vapor content of around 40 vol. %.</p>
Full article ">Scheme 1
<p>The scheme for methane catalytic oxidation reaction on the active phase of Co<sub>3</sub>O<sub>4</sub> spinel oxide.</p>
Full article ">
26 pages, 6574 KiB  
Review
Research Progress in Microporous Materials for Selective Adsorption and Separation of Methane from Low-Grade Gas
by Dongrui Su, Panpan Chen, Cunlei Li, Yongfei Yan, Ranlei Zhao, Qingyou Yue and Yupeng Qiao
Molecules 2024, 29(18), 4404; https://doi.org/10.3390/molecules29184404 - 16 Sep 2024
Viewed by 1209
Abstract
Given that methane (CH4) and nitrogen (N2) have similar properties, achieving high-purity enrichment of CH4 from nitrogen-rich low-grade gas is extremely challenging and is of great significance for sustainable development in energy and the environment. This paper reviews [...] Read more.
Given that methane (CH4) and nitrogen (N2) have similar properties, achieving high-purity enrichment of CH4 from nitrogen-rich low-grade gas is extremely challenging and is of great significance for sustainable development in energy and the environment. This paper reviews the research progress on carbon-based materials, zeolites, and MOFs as adsorbent materials for CH4/N2 separation. It focuses on the relationship between the composition, pore size, surface chemistry of the adsorbents, CH4/N2 selectivity, and CH4 adsorption capacity. The paper also highlights that controlling pore size and atomic-scale composition and optimizing these features for the best match are key directions for the development of new adsorbents. Additionally, it points out that MOFs, which combine the advantages of carbon-based adsorbents and zeolites, are likely to become the most promising adsorbent materials for efficient CH4/N2 separation. Full article
Show Figures

Figure 1

Figure 1
<p>The schematic diagram of the six-bed VPSA process [<a href="#B24-molecules-29-04404" class="html-bibr">24</a>].</p>
Full article ">Figure 2
<p>Schematic flow diagram of the eight-column VPSA process with SMB mode. Different colors represent changes in different state parameters (pressure, gas composition, etc.), and dotted lines represent the transition of the three areas [<a href="#B26-molecules-29-04404" class="html-bibr">26</a>].</p>
Full article ">Figure 3
<p>Illustration of a low-pressure foaming process by bubble growth in tar pitch with and without coal particles as additives [<a href="#B33-molecules-29-04404" class="html-bibr">33</a>].</p>
Full article ">Figure 4
<p>Schematic illustration of the synthesis of enhanced N-doped porous carbon [<a href="#B36-molecules-29-04404" class="html-bibr">36</a>].</p>
Full article ">Figure 5
<p>3D-printed activated carbon monolith [<a href="#B51-molecules-29-04404" class="html-bibr">51</a>].</p>
Full article ">Figure 6
<p>(<b>a</b>) Schematic of the synthesis strategy for PFs. (<b>b</b>–<b>d</b>) SEM image of PF-Ni-1, PF-Co-1, and PF-Mn-1. (<b>e</b>) The recording of Zeta potential and pH changes during the synthesis process. (<b>f</b>) FT-IR spectra of PF-Co-1, Asn, and BC. (<b>g</b>) Photograph of the as-obtained polymeric aerogels [<a href="#B45-molecules-29-04404" class="html-bibr">45</a>].</p>
Full article ">Figure 7
<p>Schematic of Ba-ETS-4 structural changes during heat treatment process. Titanium atoms are presented in green. Silicon atoms are presented in blue. Oxygen and chlorine are presented in red and black, respectively [<a href="#B70-molecules-29-04404" class="html-bibr">70</a>].</p>
Full article ">Figure 8
<p>N<sub>2</sub> and CH<sub>4</sub> diffusion passes in ZSM-25: (<b>a</b>) 3D view of the ZSM-25 unit cell and (<b>b</b>) four unique channels connected through eight-membered rings as the main routes for gas diffusion consisting of four double-connected cages, namely, (1) grc-opr, (2) pau-opr, (3) phi-oto, and (4) plg-oto. The door-keeping cations are highlighted [<a href="#B82-molecules-29-04404" class="html-bibr">82</a>].</p>
Full article ">Figure 9
<p>Synthesis of nanosized K-Chabazite by the seed-passaging route [<a href="#B72-molecules-29-04404" class="html-bibr">72</a>].</p>
Full article ">Figure 10
<p>(<b>a</b>) 3D framework of MIL-120Al; (<b>b</b>,<b>c</b>) molecular size of CH<sub>4</sub> and N<sub>2</sub>; (<b>d</b>) illustration of the different kinetic effects of CH<sub>4</sub> and N<sub>2</sub> through the window of MIL-120Al. Color code: C, gray; H, white; O, red; Al, cyan; N, blue [<a href="#B91-molecules-29-04404" class="html-bibr">91</a>].</p>
Full article ">Figure 11
<p>The simulated distribution of adsorption density on (<b>a</b>) CAU-10-H, (<b>b</b>) MIL-160, (<b>c</b>) Al-Fum, and (<b>d</b>) MIL-53(Al) during the adsorption process (red regions for CH<sub>4</sub>, green regions for N<sub>2</sub>) [<a href="#B92-molecules-29-04404" class="html-bibr">92</a>].</p>
Full article ">Figure 12
<p>Crystallographic structure of UiO-66 unit cell and schematic of organic linkers of UiO-66-X materials [<a href="#B98-molecules-29-04404" class="html-bibr">98</a>].</p>
Full article ">Figure 13
<p>(<b>a</b>) XRD patterns of experimental and simulated Ni-Qc-5 MOF (<b>b</b>) Thermogravimetric analysis of Ni-Qc-5 MOF [<a href="#B100-molecules-29-04404" class="html-bibr">100</a>].</p>
Full article ">Figure 14
<p>(<b>a</b>) Chemical structure of fumaric acid ligand and Zr<sub>6</sub>(μ3-O)<sub>4</sub>(μ3-OH)<sub>4</sub> cluster in MIP-203-F. (<b>b</b>) Formate-linked Zr6-oxo cluster chain along the a-axis. (<b>c</b>) Framework structure of MIP-203-F with the hydroxyl group-divided dual triangular 1D pore. (<b>d</b>) Connolly surface of MIP-203-F with a probe radius of 1.82 A viewed along the b-axis. (<b>e</b>) Van der Waals surface of MIP-203-F viewed along the a-axis [<a href="#B103-molecules-29-04404" class="html-bibr">103</a>].</p>
Full article ">
Back to TopTop